Jump to content

Hydrogen atom: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
blah blah blah
Tags: categories removed references removed
m Reverted edits by 65.74.65.84 (talk) to last version by ClueBot NG
Line 1: Line 1:
{{About|the physics of the hydrogen atom|a chemical description |hydrogen||}}
Hydrogen atom
{{Use dmy dates|date=July 2012}}
From Wikipedia, the free encyclopedia
<!-- Here is the template for this nuclide; skip past it to edit the text. -->
You have a new message from another user (last change).
{{Infobox isotope
This article is about the physics of the hydrogen atom. For a chemical description, see hydrogen.
Hydrogen atom
|image = Hydrogen 1.svg
|alternate_names = protium
Hydrogen 1.svg
|mass_number = 1
|symbol = H
|num_neutrons = 0
|num_protons = 1
|abundance = 99.985%
|mass = 1.007825
|spin = {{sfrac|1|2}}
|excess_energy = 7288.969
|error1 = 0.001
|binding_energy = 0.000
|error2 = 0.0000
}}


[[File:hydrogen atom.svg|thumb|200px|right|Depiction of a hydrogen atom showing the diameter as about twice the [[Bohr model]] radius. (Image not to scale)]]
Hydrogen atom
A '''hydrogen atom''' is an [[atom]] of the [[chemical element]] [[hydrogen]]. The [[Electric charge|electrically]] neutral atom contains a single positively charged [[proton]] and a single negatively charged [[electron]] bound to the nucleus by the [[Coulomb force]]. '''Atomic hydrogen''' constitutes [[abundance of the chemical elements|about 75%]] of the [[baryon]]ic mass of the universe.<ref>
Complete table of nuclides
{{cite web
General
|last=Palmer|first=D.
Name, symbol protium, 1H
|title=Hydrogen in the Universe
Neutrons 0
|url=http://imagine.gsfc.nasa.gov/docs/ask_astro/answers/971113i.html
Protons 1
|publisher=[[NASA]]
Nuclide data
|date=13 September 1997
Natural abundance 99.985%
|archive-url=https://web.archive.org/web/20141029152908/http://imagine.gsfc.nasa.gov/docs/ask_astro/answers/971113i.html
Isotope mass 1.007825 u
|archive-date=2014-10-29
Spin 1/2
|access-date=2017-02-23
Excess energy 7288.969± 0.001 keV
|dead-url=yes
Binding energy 0.000± 0.0000 keV
}}</ref>
Depiction of a hydrogen atom showing the diameter as about twice the Bohr model radius. (Image not to scale)


In everyday life on Earth, isolated hydrogen atoms (called "atomic hydrogen") are extremely rare. Instead, hydrogen tends to combine with other atoms in compounds, or with itself to form ordinary ([[diatomic]]) hydrogen gas, H<sub>2</sub>. "Atomic hydrogen" and "hydrogen atom" in ordinary English use have overlapping, yet distinct, meanings. For example, a water molecule contains two hydrogen atoms, but does not contain atomic hydrogen (which would refer to isolated hydrogen atoms).
A hydrogen atom is an atom of the chemical element hydrogen. The electrically neutral atom contains a single positively charged proton and a single negatively charged electron bound to the nucleus by the Coulomb force. Atomic hydrogen constitutes about 75% of the baryonic mass of the universe.[1]


Attempts to develop a theoretical understanding of the hydrogen atom have been important to the [[history of quantum mechanics]].This is the only element without a neutron
In everyday life on Earth, isolated hydrogen atoms (called "atomic hydrogen") are extremely rare. Instead, hydrogen tends to combine with other atoms in compounds, or with itself to form ordinary (diatomic) hydrogen gas, H2. "Atomic hydrogen" and "hydrogen atom" in ordinary English use have overlapping, yet distinct, meanings. For example, a water molecule contains two hydrogen atoms, but does not contain atomic hydrogen (which would refer to isolated hydrogen atoms).


==Isotopes==
Attempts to develop a theoretical understanding of the hydrogen atom have been important to the history of quantum mechanics.This is the only element without a neutron
{{Main article|Isotopes of hydrogen}}
The most [[Abundance of the chemical elements|abundant]] [[isotope]], '''hydrogen-1''', '''protium''', or '''light hydrogen''', contains no [[neutron]]s and is simply a [[proton]] and an [[electron]]. Protium is [[Proton decay|stable]] and makes up 99.985% of naturally occurring hydrogen atoms.<ref name=Housecroft>{{cite book |last1=Housecroft |first=Catherine E. |last2=Sharpe |first2=Alan G. |edition=2nd |date=2005 |title=Inorganic Chemistry |publisher=Pearson Prentice-Hall |page=237 |isbn=0130-39913-2}}</ref>


[[Deuterium]] contains one neutron and one proton. Deuterium is stable and makes up 0.0156% of naturally occurring hydrogen<ref name=Housecroft/> and is used in industrial processes like [[nuclear reactor]]s and [[Deuterium NMR|Nuclear Magnetic Resonance]].
Contents


[[Tritium]] contains two neutrons and one proton and is not stable, decaying with a [[half-life]] of 12.32 years. Because of the short half life, tritium does not exist in nature except in trace amounts.
1 Isotopes
2 Hydrogen ion
3 Theoretical analysis
3.1 Failed classical description
3.2 Bohr-Sommerfeld Model
3.3 Schrödinger equation
3.3.1 Wavefunction
3.3.2 Results of Schrödinger equation
3.3.3 Mathematical summary of eigenstates of hydrogen atom
3.3.3.1 Energy levels
3.4 Visualizing the hydrogen electron orbitals
3.4.1 Features going beyond the Schrödinger solution
4 Alternatives to the Schrödinger theory
5 See also
6 References
7 Books
8 External links


Higher isotopes of hydrogen are only created in artificial [[Particle accelerator|accelerators]] and reactors and have half lives around the order of 10<sup>−22</sup> seconds.
Isotopes
Main article: Isotopes of hydrogen


The formulas below are valid for all three isotopes of hydrogen, but slightly different values of the [[Rydberg constant]] (correction formula given below) must be used for each hydrogen isotope.
The most abundant isotope, hydrogen-1, protium, or light hydrogen, contains no neutrons and is simply a proton and an electron. Protium is stable and makes up 99.985% of naturally occurring hydrogen atoms.[2]


==Hydrogen ion==
Deuterium contains one neutron and one proton. Deuterium is stable and makes up 0.0156% of naturally occurring hydrogen[2] and is used in industrial processes like nuclear reactors and Nuclear Magnetic Resonance.
{{Main article|Hydron (chemistry)}}
Hydrogen is not found without its electron in ordinary chemistry (room temperatures and pressures), as ionized hydrogen is highly chemically reactive. When ionized hydrogen is written as "H<sup>+</sup>" as in the solvation of classical acids such as [[hydrochloric acid]], the [[hydronium ion]], [[hydrogen|H]]<sub>3</sub>[[oxygen|O]]<sup>+</sup>, is meant, not a literal ionized single hydrogen atom. In that case, the acid transfers the proton to H<sub>2</sub>O to form H<sub>3</sub>O<sup>+</sup>.


Ionized hydrogen without its electron, or free protons, are common in the [[interstellar medium]], and [[solar wind]].
Tritium contains two neutrons and one proton and is not stable, decaying with a half-life of 12.32 years. Because of the short half life, tritium does not exist in nature except in trace amounts.


==Theoretical analysis==
Higher isotopes of hydrogen are only created in artificial accelerators and reactors and have half lives around the order of 10−22 seconds.
The hydrogen atom has special significance in [[quantum mechanics]] and [[quantum field theory]] as a simple [[two-body problem]] physical system which has yielded many simple [[closed-form expression|analytical]] solutions in closed-form.


=== Failed classical description ===
The formulas below are valid for all three isotopes of hydrogen, but slightly different values of the Rydberg constant (correction formula given below) must be used for each hydrogen isotope.
[[Rutherford model|Experiments]] by [[Ernest Rutherford]] in 1909 showed the structure of the atom to be a dense, positive nucleus with a light, negative charge orbiting around it. This immediately caused problems on how such a system could be stable. [[Classical electromagnetism]] had shown that any accelerating charge radiates energy described through the [[Larmor formula]]. If the electron is assumed to orbit in a perfect circle and radiates energy continuously, the electron would rapidly spiral into the nucleus with a fall time of:<ref>{{Cite web|url = http://www.physics.princeton.edu/~mcdonald/examples/orbitdecay.pdf|title = Classical Lifetime of a Bohr Atom|date = March 7, 2005|website = |publisher = Joseph Henry Laboratories, Princeton University|last = Olsen|first = James|last2 = McDonald|first2 = Kirk}}</ref>
Hydrogen ion
Main article: Hydron (chemistry)


:<math>t_\text{fall} \approx \frac{ a_0^3}{4 r_0^2 c} \approx 1.6 \cdot 10^{-11} \text{s}</math>
Hydrogen is not found without its electron in ordinary chemistry (room temperatures and pressures), as ionized hydrogen is highly chemically reactive. When ionized hydrogen is written as "H+" as in the solvation of classical acids such as hydrochloric acid, the hydronium ion, H3O+, is meant, not a literal ionized single hydrogen atom. In that case, the acid transfers the proton to H2O to form H3O+.


Where <math>a_0</math> is the [[Bohr radius]] and <math>r_0</math> is the [[classical electron radius]]. If this were true, all atoms would instantly collapse, however atoms seem to be stable. Furthermore, the spiral inward would release a smear of electromagnetic frequencies as the orbit got smaller. Instead, atoms were observed to only emit discrete frequencies of radiation. The resolution would lie in the development of [[quantum mechanics]].
Ionized hydrogen without its electron, or free protons, are common in the interstellar medium, and solar wind.
Theoretical analysis

The hydrogen atom has special significance in quantum mechanics and quantum field theory as a simple two-body problem physical system which has yielded many simple analytical solutions in closed-form.
Failed classical description

Experiments by Ernest Rutherford in 1909 showed the structure of the atom to be a dense, positive nucleus with a light, negative charge orbiting around it. This immediately caused problems on how such a system could be stable. Classical electromagnetism had shown that any accelerating charge radiates energy described through the Larmor formula. If the electron is assumed to orbit in a perfect circle and radiates energy continuously, the electron would rapidly spiral into the nucleus with a fall time of:[3]

t fall ≈ a 0 3 4 r 0 2 c ≈ 1.6 ⋅ 10 − 11 s {\displaystyle t_{\text{fall}}\approx {\frac {a_{0}^{3}}{4r_{0}^{2}c}}\approx 1.6\cdot 10^{-11}{\text{s}}} {\displaystyle t_{\text{fall}}\approx {\frac {a_{0}^{3}}{4r_{0}^{2}c}}\approx 1.6\cdot 10^{-11}{\text{s}}}

Where a 0 {\displaystyle a_{0}} a_{0} is the Bohr radius and r 0 {\displaystyle r_{0}} r_{0} is the classical electron radius. If this were true, all atoms would instantly collapse, however atoms seem to be stable. Furthermore, the spiral inward would release a smear of electromagnetic frequencies as the orbit got smaller. Instead, atoms were observed to only emit discrete frequencies of radiation. The resolution would lie in the development of quantum mechanics.
Bohr-Sommerfeld Model
Main article: Bohr model

In 1913, Niels Bohr obtained the energy levels and spectral frequencies of the hydrogen atom after making a number of simple assumptions in order to correct the failed classical model. The assumptions included:

Electrons can only be in certain, discrete circular orbits or stationary states, thereby having a discrete set of possible radii and energies.
Electrons do not emit radiation while in one of these stationary states.
An electron can gain or lose energy by jumping from one discrete orbital to another.


=== Bohr-Sommerfeld Model ===
{{Main article|Bohr model}}In 1913, [[Niels Bohr]] obtained the energy levels and spectral frequencies of the hydrogen atom after making a number of simple assumptions in order to correct the failed classical model. The assumptions included:
# Electrons can only be in certain, discrete circular orbits or ''stationary states'', thereby having a discrete set of possible radii and energies.
# Electrons do not emit radiation while in one of these stationary states.
# An electron can gain or lose energy by jumping from one discrete orbital to another.
Bohr supposed that the electron's angular momentum is quantized with possible values:
Bohr supposed that the electron's angular momentum is quantized with possible values:


<math>L = n \hbar</math> where <math>n = 1,2,3,...</math>
L = n ℏ {\displaystyle L=n\hbar } {\displaystyle L=n\hbar } where n = 1 , 2 , 3 , . . . {\displaystyle n=1,2,3,...} {\displaystyle n=1,2,3,...}


and ℏ {\displaystyle \hbar } \hbar is Planck constant over 2 π {\displaystyle 2\pi } 2\pi . He also supposed that the centripetal force which keeps the electron in its orbit is provided by the Coulomb force, and that energy is conserved. Bohr derived the energy of each orbit of the hydrogen atom to be:[4]
and <math>\hbar</math> is [[Planck constant]] over <math>2 \pi</math>. He also supposed that the [[centripetal force]] which keeps the electron in its orbit is provided by the [[Coulomb's law|Coulomb force]], and that energy is conserved. Bohr derived the energy of each orbit of the hydrogen atom to be:<ref>{{Cite web|url = http://alpha.chem.umb.edu/chemistry/ch115/carter/files/103more/BohrEquations.pdf|title = Derivation of Bohr’s Equations for the One-electron Atom|website = |publisher = University of Massachusetts Boston|last = |first = }}</ref>
::<math>E_n = - \frac{ m_e e^4}{2 ( 4 \pi \epsilon_0)^2 \hbar^2 } \frac{1}{n^2} </math>,


where <math>m_e </math> is the [[electron mass]], <math>e </math> is the [[electron charge]], <math>\epsilon_0 </math> is the [[vacuum permittivity]], and <math>n </math> is the [[quantum number]] (now known as the [[principal quantum number]]). Bohr's predictions matched experiments measuring the [[hydrogen spectral series]] to the first order, giving more confidence to a theory that used quantized values.
E n = − m e e 4 2 ( 4 π ϵ 0 ) 2 ℏ 2 1 n 2 {\displaystyle E_{n}=-{\frac {m_{e}e^{4}}{2(4\pi \epsilon _{0})^{2}\hbar ^{2}}}{\frac {1}{n^{2}}}} {\displaystyle E_{n}=-{\frac {m_{e}e^{4}}{2(4\pi \epsilon _{0})^{2}\hbar ^{2}}}{\frac {1}{n^{2}}}},


For <math>n=1</math>, the value
where m e {\displaystyle m_{e}} {\displaystyle m_{e}} is the electron mass, e {\displaystyle e} {\displaystyle e} is the electron charge, ϵ 0 {\displaystyle \epsilon _{0}} {\displaystyle \epsilon _{0}} is the vacuum permittivity, and n {\displaystyle n} {\displaystyle n} is the quantum number (now known as the principal quantum number). Bohr's predictions matched experiments measuring the hydrogen spectral series to the first order, giving more confidence to a theory that used quantized values.
::<math>\frac{ m_e e^4}{2 ( 4 \pi \epsilon_0)^2 \hbar^2 } =\frac{m_{\text{e}} e^4}{8 h^2 \varepsilon_0^2}= 1 Ry = 13.605\;692\;53(30) \,\text{eV} </math><ref name="codata">P.J. Mohr, B.N. Taylor, and D.B. Newell (2011), "The 2010 CODATA Recommended Values of the Fundamental Physical Constants" (Web Version 6.0). This database was developed by J. Baker, M. Douma, and S. Kotochigova. Available: http://physics.nist.gov/constants. National Institute of Standards and Technology, Gaithersburg, MD 20899. [http://physics.nist.gov/cgi-bin/cuu/Value?ryd Link to R<sub>∞</sub>], [http://physics.nist.gov/cgi-bin/cuu/Value?rydhcev Link to hcR<sub>∞</sub>]</ref>
is called the Rydberg unit of energy. It is related to the [[Rydberg constant]] <math>R_\infty</math> of [[atomic physics]] by <math>1 \,\text{Ry} \equiv h c R_\infty.</math>


The exact value of the Rydberg constant assumes that the nucleus is infinitely massive with respect to the electron. For hydrogen-1, hydrogen-2 ([[deuterium]]), and hydrogen-3 ([[tritium]]) the constant must be slightly modified to use the [[reduced mass]] of the system, rather than simply the mass of the electron. However, since the nucleus is much heavier than the electron, the values are nearly the same. The Rydberg constant ''R<sub>M</sub>'' for a hydrogen atom (one electron), ''R'' is given by
For n = 1 {\displaystyle n=1} n=1, the value


<math>R_M = \frac{R_\infty}{1+m_{\text{e}}/M},</math>
m e e 4 2 ( 4 π ϵ 0 ) 2 ℏ 2 = m e e 4 8 h 2 ε 0 2 = 1 R y = 13.605 692 53 ( 30 ) eV {\displaystyle {\frac {m_{e}e^{4}}{2(4\pi \epsilon _{0})^{2}\hbar ^{2}}}={\frac {m_{\text{e}}e^{4}}{8h^{2}\varepsilon _{0}^{2}}}=1Ry=13.605\;692\;53(30)\,{\text{eV}}} {\displaystyle {\frac {m_{e}e^{4}}{2(4\pi \epsilon _{0})^{2}\hbar ^{2}}}={\frac {m_{\text{e}}e^{4}}{8h^{2}\varepsilon _{0}^{2}}}=1Ry=13.605\;692\;53(30)\,{\text{eV}}}[5]


where <math>M</math> is the mass of the atomic nucleus. For hydrogen-1, the quantity <math>m_{\text{e}}/M,</math> is about 1/1836 (i.e. the electron-to-proton mass ratio). For deuterium and tritium, the ratios are about 1/3670 and 1/5497 respectively. These figures, when added to 1 in the denominator, represent very small corrections in the value of ''R'', and thus only small corrections to all energy levels in corresponding hydrogen isotopes.
is called the Rydberg unit of energy. It is related to the Rydberg constant R ∞ {\displaystyle R_{\infty }} {\displaystyle R_{\infty }} of atomic physics by 1 Ry ≡ h c R ∞ . {\displaystyle 1\,{\text{Ry}}\equiv hcR_{\infty }.} {\displaystyle 1\,{\text{Ry}}\equiv hcR_{\infty }.}


There were still problems with Bohr's model:
The exact value of the Rydberg constant assumes that the nucleus is infinitely massive with respect to the electron. For hydrogen-1, hydrogen-2 (deuterium), and hydrogen-3 (tritium) the constant must be slightly modified to use the reduced mass of the system, rather than simply the mass of the electron. However, since the nucleus is much heavier than the electron, the values are nearly the same. The Rydberg constant RM for a hydrogen atom (one electron), R is given by
# it failed to predict other spectral details such as [[fine structure]] and [[hyperfine structure]]
# it could only predict energy levels with any accuracy for single–electron atoms (hydrogen–like atoms)
# the predicted values were only correct to <math>\alpha^2 \approx 10^{-5}</math>, where <math>\alpha</math> is the [[fine-structure constant]].


Most of these shortcomings were repaired by [[Arnold Sommerfeld|Arnold Sommerfeld's]] modification of the Bohr model. Sommerfeld introduced two additional degrees of freedom allowing an electron to move on an elliptical orbit, characterized by its [[Orbital eccentricity|eccentricity]] and [[declination]] with respect to a chosen axis. This introduces two additional quantum numbers, which correspond to the orbital [[angular momentum]] and its projection on the chosen axis. Thus the correct multiplicity of states (except for the factor 2 accounting for the yet unknown electron spin) was found. Further applying special relativity theory to the elliptic orbits, Sommerfeld succeeded in deriving the correct expression for the fine structure of hydrogen spectra (which happens to be exactly the same as in the most elaborate Dirac theory). However some observed phenomena such as the anomalous [[Zeeman effect]] remain unexplained. These issues were resolved with the full development of quantum mechanics and the [[Dirac equation]]. It is often alleged, that the [[Schrödinger equation]] is superior to the Bohr-Sommerfeld theory in describing hydrogen atom. This is however not the case, as the most results of both approaches coincide or are very close (a remarkable exception is the problem of hydrogen atom in crossed electric and magnetic fields, which cannot be solved in the framework of the Bohr-Sommerfeld theory self-consistently), and their main shortcomings result from the absence of the electron spin in both theories. It was the complete failure of the Bohr-Sommerfeld theory to explain many-electron systems (such as helium atom or hydrogen molecule) which demonstrated its inadequacy in describing quantum phenomena.
R M = R ∞ 1 + m e / M , {\displaystyle R_{M}={\frac {R_{\infty }}{1+m_{\text{e}}/M}},} R_M = \frac{R_\infty}{1+m_{\text{e}}/M},

where M {\displaystyle M} M is the mass of the atomic nucleus. For hydrogen-1, the quantity m e / M , {\displaystyle m_{\text{e}}/M,} m_{\text{e}}/M, is about 1/1836 (i.e. the electron-to-proton mass ratio). For deuterium and tritium, the ratios are about 1/3670 and 1/5497 respectively. These figures, when added to 1 in the denominator, represent very small corrections in the value of R, and thus only small corrections to all energy levels in corresponding hydrogen isotopes.

There were still problems with Bohr's model:

it failed to predict other spectral details such as fine structure and hyperfine structure
it could only predict energy levels with any accuracy for single–electron atoms (hydrogen–like atoms)
the predicted values were only correct to α 2 ≈ 10 − 5 {\displaystyle \alpha ^{2}\approx 10^{-5}} {\displaystyle \alpha ^{2}\approx 10^{-5}}, where α {\displaystyle \alpha } \alpha is the fine-structure constant.

Most of these shortcomings were repaired by Arnold Sommerfeld's modification of the Bohr model. Sommerfeld introduced two additional degrees of freedom allowing an electron to move on an elliptical orbit, characterized by its eccentricity and declination with respect to a chosen axis. This introduces two additional quantum numbers, which correspond to the orbital angular momentum and its projection on the chosen axis. Thus the correct multiplicity of states (except for the factor 2 accounting for the yet unknown electron spin) was found. Further applying special relativity theory to the elliptic orbits, Sommerfeld succeeded in deriving the correct expression for the fine structure of hydrogen spectra (which happens to be exactly the same as in the most elaborate Dirac theory). However some observed phenomena such as the anomalous Zeeman effect remain unexplained. These issues were resolved with the full development of quantum mechanics and the Dirac equation. It is often alleged, that the Schrödinger equation is superior to the Bohr-Sommerfeld theory in describing hydrogen atom. This is however not the case, as the most results of both approaches coincide or are very close (a remarkable exception is the problem of hydrogen atom in crossed electric and magnetic fields, which cannot be solved in the framework of the Bohr-Sommerfeld theory self-consistently), and their main shortcomings result from the absence of the electron spin in both theories. It was the complete failure of the Bohr-Sommerfeld theory to explain many-electron systems (such as helium atom or hydrogen molecule) which demonstrated its inadequacy in describing quantum phenomena.
Schrödinger equation


=== Schrödinger equation ===
The Schrödinger equation allows one to calculate the development of quantum systems with time and can give exact, analytical answers for the non-relativistic hydrogen atom.
The Schrödinger equation allows one to calculate the development of quantum systems with time and can give exact, analytical answers for the non-relativistic hydrogen atom.
Wavefunction


==== Wavefunction ====
The Hamiltonian of the hydrogen atom is the radial kinetic energy operator and coulomb attraction force between the positive proton and negative electron. Using the time-independent Schrödinger equation, ignoring all spin-coupling interactions and using the reduced mass μ = m e M / ( m e + M ) {\displaystyle \mu =m_{e}M/(m_{e}+M)} {\displaystyle \mu =m_{e}M/(m_{e}+M)}, the equation is written as:
The [[Hamiltonian mechanics|Hamiltonian]] of the hydrogen atom is the radial kinetic energy operator and coulomb attraction force between the positive proton and negative electron. Using the time-independent Schrödinger equation, ignoring all spin-coupling interactions and using the [[reduced mass]] <math>\mu=m_e M/(m_e + M)</math>, the equation is written as:


( − ℏ 2 2 μ ∇ 2 − e 2 4 π ϵ 0 r ) ψ ( r , θ , ϕ ) = E ψ ( r , θ , ϕ ) {\displaystyle \left(-{\frac {\hbar ^{2}}{2\mu }}\nabla ^{2}-{\frac {e^{2}}{4\pi \epsilon _{0}r}}\right)\psi (r,\theta ,\phi )=E\psi (r,\theta ,\phi )} {\displaystyle \left(-{\frac {\hbar ^{2}}{2\mu }}\nabla ^{2}-{\frac {e^{2}}{4\pi \epsilon _{0}r}}\right)\psi (r,\theta ,\phi )=E\psi (r,\theta ,\phi )}
<math>\left(- \frac{\hbar^2}{2 \mu} \nabla^2 - \frac{ e^2}{4 \pi \epsilon_0 r} \right) \psi(r,\theta, \phi) = E \psi(r, \theta, \phi)</math>


Expanding the Laplacian in spherical coordinates:
Expanding the [[Laplace operator|Laplacian]] in spherical coordinates:


− ℏ 2 2 μ [ 1 r 2 ∂ ∂ r ( r 2 ∂ ψ ∂ r ) + 1 r 2 sin ⁡ θ ∂ ∂ θ ( sin ⁡ θ ∂ ψ ∂ θ ) + 1 r 2 sin 2 ⁡ θ ∂ 2 ψ ∂ ϕ 2 ] − e 2 4 π ϵ 0 r ψ = E ψ {\displaystyle -{\frac {\hbar ^{2}}{2\mu }}\left[{\frac {1}{r^{2}}}{\frac {\partial }{\partial r}}\left(r^{2}{\frac {\partial \psi }{\partial r}}\right)+{\frac {1}{r^{2}\sin \theta }}{\frac {\partial }{\partial \theta }}\left(\sin \theta {\frac {\partial \psi }{\partial \theta }}\right)+{\frac {1}{r^{2}\sin ^{2}\theta }}{\frac {\partial ^{2}\psi }{\partial \phi ^{2}}}\right]-{\frac {e^{2}}{4\pi \epsilon _{0}r}}\psi =E\psi } {\displaystyle -{\frac {\hbar ^{2}}{2\mu }}\left[{\frac {1}{r^{2}}}{\frac {\partial }{\partial r}}\left(r^{2}{\frac {\partial \psi }{\partial r}}\right)+{\frac {1}{r^{2}\sin \theta }}{\frac {\partial }{\partial \theta }}\left(\sin \theta {\frac {\partial \psi }{\partial \theta }}\right)+{\frac {1}{r^{2}\sin ^{2}\theta }}{\frac {\partial ^{2}\psi }{\partial \phi ^{2}}}\right]-{\frac {e^{2}}{4\pi \epsilon _{0}r}}\psi =E\psi }
<math>-\frac{\hbar^2}{2 \mu} \left[ \frac{1}{r^2} \frac{\partial }{\partial r} \left( r^2 \frac{ \partial \psi}{\partial r}\right) + \frac{1}{r^2 \sin \theta} \frac{\partial }{\partial \theta} \left( \sin \theta \frac{\partial \psi}{\partial \theta}\right) + \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2 \psi}{\partial \phi^2} \right] - \frac{e^2}{ 4 \pi \epsilon_0 r} \psi= E \psi</math>

This is a separable, partial differential equation which can be solved in terms of special functions. The normalized position wavefunctions, given in spherical coordinates are:

ψ n ℓ m ( r , ϑ , φ ) = ( 2 n a 0 ∗ ) 3 ( n − ℓ − 1 ) ! 2 n ( ( n + ℓ ) ! ) 3 e − ρ / 2 ρ ℓ L n − ℓ − 1 2 ℓ + 1 ( ρ ) Y ℓ m ( ϑ , φ ) {\displaystyle \psi _{n\ell m}(r,\vartheta ,\varphi )={\sqrt {{\left({\frac {2}{na_{0}^{*}}}\right)}^{3}{\frac {(n-\ell -1)!}{2n\left((n+\ell )!\right)^{3}}}}}e^{-\rho /2}\rho ^{\ell }L_{n-\ell -1}^{2\ell +1}(\rho )Y_{\ell }^{m}(\vartheta ,\varphi )} {\displaystyle \psi _{n\ell m}(r,\vartheta ,\varphi )={\sqrt {{\left({\frac {2}{na_{0}^{*}}}\right)}^{3}{\frac {(n-\ell -1)!}{2n\left((n+\ell )!\right)^{3}}}}}e^{-\rho /2}\rho ^{\ell }L_{n-\ell -1}^{2\ell +1}(\rho )Y_{\ell }^{m}(\vartheta ,\varphi )}

3D illustration of the eigenstate ψ 4 , 3 , 1 {\displaystyle \psi _{4,3,1}} \psi_{4,3,1}. Electrons in this state are 45% likely to be found within the solid body shown.


This is a [[Separation of variables|separable]], [[partial differential equation]] which can be solved in terms of special functions. The normalized position [[wavefunction]]s, given in [[spherical coordinates]] are:
::<math> \psi_{n\ell m}(r,\vartheta,\varphi) = \sqrt {{\left ( \frac{2}{n a^*_0} \right )}^3 \frac{(n-\ell-1)!}{2n\left((n+\ell)!\right)^3}} e^{- \rho / 2} \rho^{\ell} L_{n-\ell-1}^{2\ell+1}(\rho) Y_{\ell}^{m}(\vartheta, \varphi ) </math>
[[Image:Hydrogen eigenstate n4 l3 m1.png|thumb|right|3D illustration of the eigenstate <math>\psi_{4,3,1}</math>. Electrons in this state are 45% likely to be found within the solid body shown.]]
where:
where:
::<math> \rho = {2r \over {n a^*_0}} </math>,
::<math> a^*_0 </math> is the [[reduced Bohr radius]], <math>a^*_0={{4\pi\epsilon_0\hbar^2}\over{\mu e^2}}</math>,
::<math> L_{n-\ell-1}^{2\ell+1}(\rho) </math> is a [[Laguerre polynomial#Generalized Laguerre polynomials|generalized Laguerre polynomial]] of degree {{nowrap|''n'' − ''ℓ'' − 1}}, and
::<math> Y_{\ell}^{m}(\vartheta, \varphi ) \,</math> is a [[spherical harmonic]] function of degree ''ℓ'' and order ''m''. Note that the generalized Laguerre polynomials are defined differently by different authors. The usage here is consistent with the definitions used by Messiah,<ref>{{cite book|last=Messiah|first=Albert|title=Quantum Mechanics|date=1999|publisher=Dover|location=New York|isbn=0-486-40924-4|pages=1136}}</ref> and Mathematica.<ref>[http://reference.wolfram.com/mathematica/ref/LaguerreL.html LaguerreL]. Wolfram Mathematica page</ref> In other places, the Laguerre polynomial includes a factor of <math>(n+\ell)!</math>,<ref>Griffiths, p. 152</ref> or the generalized Laguerre polynomial appearing in the hydrogen wave function is <math> L_{n+\ell}^{2\ell+1}(\rho)</math> instead.<ref>{{cite book|last=Condon and Shortley|title=The Theory of Atomic Spectra|date=1963|publisher=Cambridge|location=London|pages=441}}</ref>


The quantum numbers can take the following values:
ρ = 2 r n a 0 ∗ {\displaystyle \rho ={2r \over {na_{0}^{*}}}} {\displaystyle \rho ={2r \over {na_{0}^{*}}}},
::<math> n=1,2,3,\ldots </math>
a 0 ∗ {\displaystyle a_{0}^{*}} {\displaystyle a_{0}^{*}} is the reduced Bohr radius, a 0 ∗ = 4 π ϵ 0 ℏ 2 μ e 2 {\displaystyle a_{0}^{*}={{4\pi \epsilon _{0}\hbar ^{2}} \over {\mu e^{2}}}} {\displaystyle a_{0}^{*}={{4\pi \epsilon _{0}\hbar ^{2}} \over {\mu e^{2}}}},
::<math>\ell=0,1,2,\ldots,n-1</math>
L n − ℓ − 1 2 ℓ + 1 ( ρ ) {\displaystyle L_{n-\ell -1}^{2\ell +1}(\rho )} L_{n-\ell-1}^{2\ell+1}(\rho) is a generalized Laguerre polynomial of degree n − ℓ − 1, and
::<math>m=-\ell,\ldots,\ell.</math>
Y ℓ m ( ϑ , φ ) {\displaystyle Y_{\ell }^{m}(\vartheta ,\varphi )\,} Y_{\ell}^{m}(\vartheta, \varphi ) \, is a spherical harmonic function of degree ℓ and order m. Note that the generalized Laguerre polynomials are defined differently by different authors. The usage here is consistent with the definitions used by Messiah,[6] and Mathematica.[7] In other places, the Laguerre polynomial includes a factor of ( n + ℓ ) ! {\displaystyle (n+\ell )!} (n+\ell)!,[8] or the generalized Laguerre polynomial appearing in the hydrogen wave function is L n + ℓ 2 ℓ + 1 ( ρ ) {\displaystyle L_{n+\ell }^{2\ell +1}(\rho )} L_{n+\ell}^{2\ell+1}(\rho) instead.[9]


Additionally, these wavefunctions are ''normalized'' (i.e., the integral of their modulus square equals 1) and [[Orthogonal functions|orthogonal]]:
The quantum numbers can take the following values:
::<math>\int_0^{\infty} r^2 dr\int_0^{\pi} \sin \vartheta d\vartheta \int_0^{2 \pi} d\varphi\; \psi^*_{n\ell m}(r,\vartheta,\varphi)\psi_{n'\ell'm'}(r,\vartheta,\varphi)=\langle n,\ell, m | n', \ell', m' \rangle = \delta_{nn'} \delta_{\ell\ell'} \delta_{mm'},</math>

where <math>| n, \ell, m \rangle</math> is the state represented by the wavefunction <math> \psi_{n\ell m} </math> in [[Dirac notation]], and <math> \delta </math> is the [[Kronecker delta]] function.<ref>Griffiths, Ch. 4 p. 89</ref>
n = 1 , 2 , 3 , … {\displaystyle n=1,2,3,\ldots } n=1,2,3,\ldots
ℓ = 0 , 1 , 2 , … , n − 1 {\displaystyle \ell =0,1,2,\ldots ,n-1} \ell=0,1,2,\ldots,n-1
m = − ℓ , … , ℓ . {\displaystyle m=-\ell ,\ldots ,\ell .} m=-\ell,\ldots,\ell.

Additionally, these wavefunctions are normalized (i.e., the integral of their modulus square equals 1) and orthogonal:

∫ 0 ∞ r 2 d r ∫ 0 π sin ⁡ ϑ d ϑ ∫ 0 2 π d φ ψ n ℓ m ∗ ( r , ϑ , φ ) ψ n ′ ℓ ′ m ′ ( r , ϑ , φ ) = ⟨ n , ℓ , m | n ′ , ℓ ′ , m ′ ⟩ = δ n n ′ δ ℓ ℓ ′ δ m m ′ , {\displaystyle \int _{0}^{\infty }r^{2}dr\int _{0}^{\pi }\sin \vartheta d\vartheta \int _{0}^{2\pi }d\varphi \;\psi _{n\ell m}^{*}(r,\vartheta ,\varphi )\psi _{n'\ell 'm'}(r,\vartheta ,\varphi )=\langle n,\ell ,m|n',\ell ',m'\rangle =\delta _{nn'}\delta _{\ell \ell '}\delta _{mm'},} \int_0^{\infty} r^2 dr\int_0^{\pi} \sin \vartheta d\vartheta \int_0^{2 \pi} d\varphi\; \psi^*_{n\ell m}(r,\vartheta,\varphi)\psi_{n'\ell'm'}(r,\vartheta,\varphi)=\langle n,\ell, m | n', \ell', m' \rangle = \delta_{nn'} \delta_{\ell\ell'} \delta_{mm'},

where | n , ℓ , m ⟩ {\displaystyle |n,\ell ,m\rangle } | n, \ell, m \rangle is the state represented by the wavefunction ψ n ℓ m {\displaystyle \psi _{n\ell m}} \psi_{n\ell m} in Dirac notation, and δ {\displaystyle \delta } \delta is the Kronecker delta function.[10]


The wavefunctions in momentum space are related to the wavefunctions in position space through a Fourier transform
The wavefunctions in momentum space are related to the wavefunctions in position space through a Fourier transform


ϕ ( p , ϑ p , φ p ) = ( 2 π ℏ ) − 3 / 2 ∫ e − i p → ⋅ r → / ℏ ψ ( r , ϑ , φ ) d V , {\displaystyle \phi (p,\vartheta _{p},\varphi _{p})=(2\pi \hbar )^{-3/2}\int e^{-i{\vec {p}}\cdot {\vec {r}}/\hbar }\psi (r,\vartheta ,\varphi )dV,} \phi (p,\vartheta _{p},\varphi _{p})=(2\pi \hbar )^{{-3/2}}\int e^{{-i{\vec {p}}\cdot {\vec {r}}/\hbar }}\psi (r,\vartheta ,\varphi )dV,
::<math> \phi(p, \vartheta_p, \varphi_p) = (2\pi\hbar)^{-3/2} \int e^{-i \vec{p} \cdot \vec{r} / \hbar} \psi(r,\vartheta,\varphi) dV, </math>


which, for the bound states, results in [11]
which, for the bound states, results in <ref>{{cite book
| first=B. H.
| last=Bransden
|author2=Joachain, C. J.
| title=Physics of Atoms and Molecules
| publisher=[[Longman]]
| date=1983
| isbn=0-582-44401-2| page = Appendix 5
}}</ref>


ϕ ( p , ϑ p , φ p ) = 2 π ( n − l − 1 ) ! ( n + l ) ! n 2 2 2 l + 2 l ! n l p l ( n 2 p 2 + 1 ) l + 2 C n − l − 1 l + 1 ( n 2 p 2 − 1 n 2 p 2 + 1 ) Y l m ( ϑ p , φ p ) , {\displaystyle \phi (p,\vartheta _{p},\varphi _{p})={\sqrt {{\frac {2}{\pi }}{\frac {(n-l-1)!}{(n+l)!}}}}n^{2}2^{2l+2}l!{\frac {n^{l}p^{l}}{(n^{2}p^{2}+1)^{l+2}}}C_{n-l-1}^{l+1}\left({\frac {n^{2}p^{2}-1}{n^{2}p^{2}+1}}\right)Y_{l}^{m}({\vartheta _{p},\varphi _{p}}),} \phi (p,\vartheta _{p},\varphi _{p})={\sqrt {{\frac {2}{\pi }}{\frac {(n-l-1)!}{(n+l)!}}}}n^{2}2^{{2l+2}}l!{\frac {n^{l}p^{l}}{(n^{2}p^{2}+1)^{{l+2}}}}C_{{n-l-1}}^{{l+1}}\left({\frac {n^{2}p^{2}-1}{n^{2}p^{2}+1}}\right)Y_{l}^{m}({\vartheta _{p},\varphi _{p}}),
::<math> \phi(p, \vartheta_p, \varphi_p) = \sqrt{\frac{2}{\pi} \frac{(n-l-1)!}{(n+l)!}} n^2 2^{2l+2} l! \frac{n^l p^l}{(n^2 p^2 + 1)^{l+2}} C_{n-l-1}^{l+1}\left(\frac{n^2 p^2 - 1}{n^2 p^2 + 1}\right) Y_l^m({\vartheta_p, \varphi_p}), </math>


where C N α ( x ) {\displaystyle C_{N}^{\alpha }(x)} C_{N}^{\alpha }(x) denotes a Gegenbauer polynomial and p {\displaystyle p} p is in units of ℏ / a 0 ∗ {\displaystyle \hbar /a_{0}^{*}} {\displaystyle \hbar /a_{0}^{*}}.
where <math> C_N^\alpha(x) </math> denotes a [[Gegenbauer polynomial]] and <math> p </math> is in units of <math> \hbar/a^*_0 </math>.


The solutions to the Schrödinger equation for hydrogen are analytical, giving a simple expression for the hydrogen energy levels and thus the frequencies of the hydrogen spectral lines and fully reproduced the Bohr model and went beyond it. It also yields two other quantum numbers and the shape of the electron's wave function ("orbital") for the various possible quantum-mechanical states, thus explaining the anisotropic character of atomic bonds.
The solutions to the Schrödinger equation for hydrogen are [[analytical expression|analytical]], giving a simple expression for the hydrogen [[energy levels]] and thus the frequencies of the hydrogen [[spectral line]]s and fully reproduced the Bohr model and went beyond it. It also yields two other quantum numbers and the shape of the electron's wave function ("orbital") for the various possible quantum-mechanical states, thus explaining the [[anisotropic]] character of atomic bonds.


The Schrödinger equation also applies to more complicated atoms and molecules. When there is more than one electron or nucleus the solution is not analytical and either computer calculations are necessary or simplifying assumptions must be made.
The Schrödinger equation also applies to more complicated atoms and [[molecule]]s. When there is more than one electron or nucleus the solution is not analytical and either computer calculations are necessary or simplifying assumptions must be made.


Since the Schrödinger equation is only valid for non-relativistic quantum mechanics, the solutions it yields for the hydrogen atom are not entirely correct. The Dirac equation of relativistic quantum theory improves these solutions (see below).
Since the Schrödinger equation is only valid for non-relativistic quantum mechanics, the solutions it yields for the hydrogen atom are not entirely correct. The [[Dirac equation]] of relativistic quantum theory improves these solutions (see below).
Results of Schrödinger equation


==== Results of Schrödinger equation ====
The solution of the Schrödinger equation (wave equation) for the hydrogen atom uses the fact that the Coulomb potential produced by the nucleus is isotropic (it is radially symmetric in space and only depends on the distance to the nucleus). Although the resulting energy eigenfunctions (the orbitals) are not necessarily isotropic themselves, their dependence on the angular coordinates follows completely generally from this isotropy of the underlying potential: the eigenstates of the Hamiltonian (that is, the energy eigenstates) can be chosen as simultaneous eigenstates of the angular momentum operator. This corresponds to the fact that angular momentum is conserved in the orbital motion of the electron around the nucleus. Therefore, the energy eigenstates may be classified by two angular momentum quantum numbers, ℓ and m (both are integers). The angular momentum quantum number ℓ = 0, 1, 2, ... determines the magnitude of the angular momentum. The magnetic quantum number m = −ℓ, ..., +ℓ determines the projection of the angular momentum on the (arbitrarily chosen) z-axis.
The solution of the Schrödinger equation (wave equation) for the hydrogen atom uses the fact that the [[Coulomb's law|Coulomb potential]] produced by the nucleus is [[isotropic]] (it is radially symmetric in space and only depends on the distance to the nucleus). Although the resulting [[energy eigenfunctions]] (the ''orbitals'') are not necessarily isotropic themselves, their dependence on the [[Spherical coordinate system|angular coordinates]] follows completely generally from this isotropy of the underlying potential: the [[eigenstates]] of the [[Hamiltonian (quantum mechanics)|Hamiltonian]] (that is, the energy eigenstates) can be chosen as simultaneous eigenstates of the [[angular momentum operator]]. This corresponds to the fact that angular momentum is conserved in the [[orbital motion (quantum)|orbital motion]] of the electron around the nucleus. Therefore, the energy eigenstates may be classified by two angular momentum [[quantum number]]s, ''ℓ'' and ''m'' (both are integers). The angular momentum quantum number {{nowrap|''ℓ'' {{=}} 0, 1, 2, ...}} determines the magnitude of the angular momentum. The magnetic quantum number {{nowrap|''m'' {{=}} −''ℓ'', ..., +''ℓ''}} determines the projection of the angular momentum on the (arbitrarily chosen) ''z''-axis.


In addition to mathematical expressions for total angular momentum and angular momentum projection of wavefunctions, an expression for the radial dependence of the wave functions must be found. It is only here that the details of the 1/r Coulomb potential enter (leading to Laguerre polynomials in r). This leads to a third quantum number, the principal quantum number n = 1, 2, 3, .... The principal quantum number in hydrogen is related to the atom's total energy.
In addition to mathematical expressions for total angular momentum and angular momentum projection of wavefunctions, an expression for the radial dependence of the wave functions must be found. It is only here that the details of the 1/''r'' Coulomb potential enter (leading to [[Laguerre polynomials]] in ''r''). This leads to a third quantum number, the principal quantum number {{nowrap|''n'' {{=}} 1, 2, 3, ...}}. The principal quantum number in hydrogen is related to the atom's total energy.


Note that the maximum value of the angular momentum quantum number is limited by the principal quantum number: it can run only up to n − 1, i.e. ℓ = 0, 1, ..., n − 1.
Note that the maximum value of the angular momentum quantum number is limited by the principal quantum number: it can run only up to ''n'' − 1, i.e. {{nowrap|'''' {{=}} 0, 1, ..., ''n'' − 1}}.


Due to angular momentum conservation, states of the same ℓ but different m have the same energy (this holds for all problems with rotational symmetry). In addition, for the hydrogen atom, states of the same n but different ℓ are also degenerate (i.e. they have the same energy). However, this is a specific property of hydrogen and is no longer true for more complicated atoms which have an (effective) potential differing from the form 1/r (due to the presence of the inner electrons shielding the nucleus potential).
Due to angular momentum conservation, states of the same '''' but different ''m'' have the same energy (this holds for all problems with [[rotational symmetry]]). In addition, for the hydrogen atom, states of the same ''n'' but different '''' are also [[degenerate energy levels|degenerate]] (i.e. they have the same energy). However, this is a specific property of hydrogen and is no longer true for more complicated atoms which have an (effective) potential differing from the form 1/''r'' (due to the presence of the inner electrons shielding the nucleus potential).


Taking into account the spin of the electron adds a last quantum number, the projection of the electron's spin angular momentum along the z-axis, which can take on two values. Therefore, any eigenstate of the electron in the hydrogen atom is described fully by four quantum numbers. According to the usual rules of quantum mechanics, the actual state of the electron may be any superposition of these states. This explains also why the choice of z-axis for the directional quantization of the angular momentum vector is immaterial: an orbital of given ℓ and m′ obtained for another preferred axis z′ can always be represented as a suitable superposition of the various states of different m (but same l) that have been obtained for z.
Taking into account the [[spin (physics)|spin]] of the electron adds a last quantum number, the projection of the electron's spin angular momentum along the ''z''-axis, which can take on two values. Therefore, any [[eigenstate]] of the electron in the hydrogen atom is described fully by four quantum numbers. According to the usual rules of quantum mechanics, the actual state of the electron may be any [[quantum superposition|superposition]] of these states. This explains also why the choice of ''z''-axis for the directional [[quantization (physics)|quantization]] of the angular momentum vector is immaterial: an orbital of given '''' and ''m''&prime; obtained for another preferred axis ''z''&prime; can always be represented as a suitable superposition of the various states of different ''m'' (but same ''l'') that have been obtained for ''z''.
Mathematical summary of eigenstates of hydrogen atom
Main article: Hydrogen-like atom


==== Mathematical summary of eigenstates of hydrogen atom ====
In 1928, Paul Dirac found an equation that was fully compatible with Special Relativity, and (as a consequence) made the wave function a 4-component "Dirac spinor" including "up" and "down" spin components, with both positive and "negative" energy (or matter and antimatter). The solution to this equation gave the following results, more accurate than the Schrödinger solution.
{{main article|Hydrogen-like atom}}
Energy levels


In 1928, [[Paul Dirac]] found [[Dirac equation|an equation]] that was fully compatible with [[Special Relativity]], and (as a consequence) made the wave function a 4-component "[[Dirac spinor]]" including "up" and "down" spin components, with both positive and "negative" energy (or matter and antimatter). The solution to this equation gave the following results, more accurate than the Schrödinger solution.
The energy levels of hydrogen, including fine structure (excluding Lamb shift and hyperfine structure), are given by the Sommerfeld fine structure expression:[12]


=====Energy levels=====
E j n = − μ c 2 [ 1 − ( 1 + [ α n − j − 1 2 + ( j + 1 2 ) 2 − α 2 ] 2 ) − 1 / 2 ] ≈ − μ c 2 α 2 2 n 2 [ 1 + α 2 n 2 ( n j + 1 2 − 3 4 ) ] , {\displaystyle {\begin{array}{rl}E_{j\,n}&=-\mu c^{2}\left[1-\left(1+\left[{\dfrac {\alpha }{n-j-{\frac {1}{2}}+{\sqrt {\left(j+{\frac {1}{2}}\right)^{2}-\alpha ^{2}}}}}\right]^{2}\right)^{-1/2}\right]\\&\approx -{\dfrac {\mu c^{2}\alpha ^{2}}{2n^{2}}}\left[1+{\dfrac {\alpha ^{2}}{n^{2}}}\left({\dfrac {n}{j+{\frac {1}{2}}}}-{\dfrac {3}{4}}\right)\right],\end{array}}} {\displaystyle {\begin{array}{rl}E_{j\,n}&=-\mu c^{2}\left[1-\left(1+\left[{\dfrac {\alpha }{n-j-{\frac {1}{2}}+{\sqrt {\left(j+{\frac {1}{2}}\right)^{2}-\alpha ^{2}}}}}\right]^{2}\right)^{-1/2}\right]\\&\approx -{\dfrac {\mu c^{2}\alpha ^{2}}{2n^{2}}}\left[1+{\dfrac {\alpha ^{2}}{n^{2}}}\left({\dfrac {n}{j+{\frac {1}{2}}}}-{\dfrac {3}{4}}\right)\right],\end{array}}}
The energy levels of hydrogen, including [[fine structure]] (excluding [[Lamb shift]] and [[hyperfine structure]]), are given by the [[Fine-structure constant|Sommerfeld fine structure]] expression:<ref name="Sommerfeld">{{cite book|first=Arnold |last= Sommerfeld|title=Atombau und Spektrallinien'| publisher=Friedrich Vieweg und Sohn|location=Braunschweig|year=1919|isbn=3-87144-484-7}} [https://archive.org/stream/atombauundspekt00sommgoog German] [https://archive.org/details/AtomicStructureAndSpectralLines English]</ref>
::<math>\begin{array}{rl} E_{j\,n} & = -\mu c^2\left[1-\left(1+\left[\dfrac{\alpha}{n-j-\frac{1}{2}+\sqrt{\left(j+\frac{1}{2}\right)^2-\alpha^2}}\right]^2\right)^{-1/2}\right] \\ & \approx -\dfrac{\mu c^2\alpha^2}{2n^2} \left[1 + \dfrac{\alpha^2}{n^2}\left(\dfrac{n}{j+\frac{1}{2}} - \dfrac{3}{4} \right) \right] , \end{array}</math>


where α is the fine-structure constant and j is the "total angular momentum" quantum number, which is equal to |ℓ ± 1/2| depending on the direction of the electron spin. This formula represents a small correction to the energy obtained by Bohr and Schrödinger as given above. The factor in square brackets in the last expression is nearly one; the extra term arises from relativistic effects (for details, see #Features going beyond the Schrödinger solution). It is worth noting that this expression was first obtained by A. Sommerfeld in 1916 based on the relativistic version of the old Bohr theory. Sommerfeld has however used different notation for the quantum numbers.
where ''α'' is the [[fine-structure constant]] and ''j'' is the "total angular momentum" [[quantum number]], which is equal to |'''' ± {{sfrac|1|2}}| depending on the direction of the electron spin. This formula represents a small correction to the energy obtained by Bohr and Schrödinger as given above. The factor in square brackets in the last expression is nearly one; the extra term arises from relativistic effects (for details, see [[#Features going beyond the Schrödinger solution]]). It is worth noting that this expression was first obtained by [[Arnold Sommerfeld|A. Sommerfeld]] in 1916 based on the relativistic version of the [[Old quantum theory|old Bohr theory]]. Sommerfeld has however used different notation for the quantum numbers.
Visualizing the hydrogen electron orbitals
Main article: Atomic orbital
Probability densities through the xz-plane for the electron at different quantum numbers (ℓ, across top; n, down side; m = 0)


=== Visualizing the hydrogen electron orbitals ===
The image to the right shows the first few hydrogen atom orbitals (energy eigenfunctions). These are cross-sections of the probability density that are color-coded (black represents zero density and white represents the highest density). The angular momentum (orbital) quantum number ℓ is denoted in each column, using the usual spectroscopic letter code (s means ℓ = 0, p means ℓ = 1, d means ℓ = 2). The main (principal) quantum number n (= 1, 2, 3, ...) is marked to the right of each row. For all pictures the magnetic quantum number m has been set to 0, and the cross-sectional plane is the xz-plane (z is the vertical axis). The probability density in three-dimensional space is obtained by rotating the one shown here around the z-axis.
{{main article|Atomic orbital}}
[[File:HAtomOrbitals.png|frame|Probability densities through the ''xz''-plane for the electron at different quantum numbers (''ℓ'', across top; ''n'', down side; ''m'' = 0)]]
The image to the right shows the first few hydrogen atom orbitals (energy eigenfunctions). These are cross-sections of the [[probability amplitude|probability density]] that are color-coded (black represents zero density and white represents the highest density). The angular momentum (orbital) quantum number ''ℓ'' is denoted in each column, using the usual spectroscopic letter code (''s'' means ''ℓ''&nbsp;=&nbsp;0, ''p'' means ''ℓ''&nbsp;=&nbsp;1, ''d'' means ''ℓ''&nbsp;=&nbsp;2). The main (principal) quantum number ''n'' (= 1, 2, 3, ...) is marked to the right of each row. For all pictures the magnetic quantum number ''m'' has been set to 0, and the cross-sectional plane is the ''xz''-plane (''z'' is the vertical axis). The probability density in three-dimensional space is obtained by rotating the one shown here around the ''z''-axis.


The "ground state", i.e. the state of lowest energy, in which the electron is usually found, is the first one, the 1s state (principal quantum level n = 1, ℓ = 0).
The "[[ground state]]", i.e. the state of lowest energy, in which the electron is usually found, is the first one, the 1''s'' state ([[principal quantum level]] ''n'' = 1, '''' = 0).


Black lines occur in each but the first orbital: these are the nodes of the wavefunction, i.e. where the probability density is zero. (More precisely, the nodes are spherical harmonics that appear as a result of solving Schrödinger equation in polar coordinates.)
Black lines occur in each but the first orbital: these are the nodes of the wavefunction, i.e. where the probability density is zero. (More precisely, the nodes are [[spherical harmonics]] that appear as a result of solving [[Schrödinger equation]] in polar coordinates.)


The quantum numbers determine the layout of these nodes.[13] There are:
The [[quantum number]]s determine the layout of these nodes.<ref>[http://www.physics.byu.edu/faculty/durfee/courses/Summer2009/physics222/AtomicQuantumNumbers.pdf Summary of atomic quantum numbers]. Lecture notes. 28 July 2006</ref> There are:
* <math>n-1</math> total nodes,

* <math>l</math> of which are angular nodes:
n − 1 {\displaystyle n-1} n-1 total nodes,
** <math>m</math> angular nodes go around the <math>\phi</math> axis (in the xy plane). <small>(The figure above does not show these nodes since it plots cross-sections through the xz-plane.)</small>
l {\displaystyle l} l of which are angular nodes:
** <math>l-m</math> (the remaining angular nodes) occur on the <math>\theta</math> (vertical) axis.
m {\displaystyle m} m angular nodes go around the ϕ {\displaystyle \phi } \phi axis (in the xy plane). (The figure above does not show these nodes since it plots cross-sections through the xz-plane.)
l − m {\displaystyle l-m} l-m (the remaining angular nodes) occur on the θ {\displaystyle \theta } \theta (vertical) axis.
* <math>n - l - 1</math> (the remaining non-angular nodes) are radial nodes.
n − l − 1 {\displaystyle n-l-1} n - l - 1 (the remaining non-angular nodes) are radial nodes.

Features going beyond the Schrödinger solution


==== Features going beyond the Schrödinger solution ====
There are several important effects that are neglected by the Schrödinger equation and which are responsible for certain small but measurable deviations of the real spectral lines from the predicted ones:
There are several important effects that are neglected by the Schrödinger equation and which are responsible for certain small but measurable deviations of the real spectral lines from the predicted ones:


Although the mean speed of the electron in hydrogen is only 1/137th of the speed of light, many modern experiments are sufficiently precise that a complete theoretical explanation requires a fully relativistic treatment of the problem. A relativistic treatment results in a momentum increase of about 1 part in 37,000 for the electron. Since the electron's wavelength is determined by its momentum, orbitals containing higher speed electrons show contraction due to smaller wavelengths.
* Although the mean speed of the electron in hydrogen is only 1/137th of the [[speed of light]], many modern experiments are sufficiently precise that a complete theoretical explanation requires a fully relativistic treatment of the problem. A relativistic treatment results in a momentum increase of about 1 part in 37,000 for the electron. Since the electron's wavelength is determined by its momentum, orbitals containing higher speed electrons show contraction due to smaller wavelengths.
Even when there is no external magnetic field, in the inertial frame of the moving electron, the electromagnetic field of the nucleus has a magnetic component. The spin of the electron has an associated magnetic moment which interacts with this magnetic field. This effect is also explained by special relativity, and it leads to the so-called spin-orbit coupling, i.e., an interaction between the electron's orbital motion around the nucleus, and its spin.
* Even when there is no external [[magnetic field]], in the [[inertial frame]] of the moving electron, the electromagnetic field of the nucleus has a magnetic component. The spin of the electron has an associated [[magnetic moment]] which interacts with this magnetic field. This effect is also explained by special relativity, and it leads to the so-called ''[[spin-orbit coupling]]'', i.e., an interaction between the [[electron]]'s [[orbital motion (quantum)|orbital motion]] around the nucleus, and its [[spin (physics)|spin]].


Both of these features (and more) are incorporated in the relativistic Dirac equation, with predictions that come still closer to experiment. Again the Dirac equation may be solved analytically in the special case of a two-body system, such as the hydrogen atom. The resulting solution quantum states now must be classified by the total angular momentum number j (arising through the coupling between electron spin and orbital angular momentum). States of the same j and the same n are still degenerate. Thus, direct analytical solution of Dirac equation predicts 2S(1/2) and 2P(1/2) levels of Hydrogen to have exactly the same energy, which is in a contradiction with observations (Lamb-Retherford experiment).
Both of these features (and more) are incorporated in the relativistic [[Dirac equation]], with predictions that come still closer to experiment. Again the Dirac equation may be solved analytically in the special case of a two-body system, such as the hydrogen atom. The resulting solution quantum states now must be classified by the [[Total angular momentum quantum number|total angular momentum number]] ''j'' (arising through the coupling between [[electron spin]] and [[angular momentum operator|orbital angular momentum]]). States of the same ''j'' and the same ''n'' are still degenerate. Thus, direct analytical solution of [[Dirac equation]] predicts 2S({{sfrac|1|2}}) and 2P({{sfrac|1|2}}) levels of Hydrogen to have exactly the same energy, which is in a contradiction with observations ([[Lamb shift|Lamb-Retherford experiment]]).


There are always vacuum fluctuations of the electromagnetic field, according to quantum mechanics. Due to such fluctuations degeneracy between states of the same j but different l is lifted, giving them slightly different energies. This has been demonstrated in the famous Lamb-Retherford experiment and was the starting point for the development of the theory of Quantum electrodynamics (which is able to deal with these vacuum fluctuations and employs the famous Feynman diagrams for approximations using perturbation theory). This effect is now called Lamb shift.
* There are always [[quantum fluctuation|vacuum fluctuation]]s of the [[electromagnetic field]], according to quantum mechanics. Due to such fluctuations degeneracy between states of the same j but different l is lifted, giving them slightly different energies. This has been demonstrated in the famous [[Lamb shift|Lamb-Retherford experiment]] and was the starting point for the development of the theory of [[Quantum electrodynamics]] (which is able to deal with these vacuum fluctuations and employs the famous [[Feynman diagram]]s for approximations using [[perturbation theory (quantum mechanics)|perturbation theory]]). This effect is now called [[Lamb shift]].


For these developments, it was essential that the solution of the Dirac equation for the hydrogen atom could be worked out exactly, such that any experimentally observed deviation had to be taken seriously as a signal of failure of the theory.
For these developments, it was essential that the solution of the Dirac equation for the hydrogen atom could be worked out exactly, such that any experimentally observed deviation had to be taken seriously as a signal of failure of the theory.
Alternatives to the Schrödinger theory

In the language of Heisenberg's matrix mechanics, the hydrogen atom was first solved by Wolfgang Pauli[14] using a rotational symmetry in four dimensions [O(4)-symmetry] generated by the angular momentum and the Laplace–Runge–Lenz vector. By extending the symmetry group O(4) to the dynamical group O(4,2), the entire spectrum and all transitions were embedded in a single irreducible group representation.[15]

In 1979 the (non relativistic) hydrogen atom was solved for the first time within Feynman's path integral formulation of quantum mechanics.[16][17] This work greatly extended the range of applicability of Feynman's method.
See also

Book: Hydrogen

Antihydrogen
Atomic orbital
Balmer series
Helium atom
Lithium atom
Hydrogen Molecular Ion


Proton decay
Quantum chemistry
Quantum state
Theoretical and experimental justification for the Schrödinger equation
Trihydrogen cation
List of quantum-mechanical systems with analytical solutions

References


== Alternatives to the Schrödinger theory ==
Palmer, D. (13 September 1997). "Hydrogen in the Universe". NASA. Archived from the original on 2014-10-29. Retrieved 2017-02-23.
In the language of Heisenberg's matrix mechanics, the hydrogen atom was first solved by [[Wolfgang Pauli]]<ref name="pauli_1926">
Housecroft, Catherine E.; Sharpe, Alan G. (2005). Inorganic Chemistry (2nd ed.). Pearson Prentice-Hall. p. 237. ISBN 0130-39913-2.
{{cite journal
Olsen, James; McDonald, Kirk (March 7, 2005). "Classical Lifetime of a Bohr Atom" (PDF). Joseph Henry Laboratories, Princeton University.
| last = Pauli | first = W | authorlink = Wolfgang Pauli
"Derivation of Bohr's Equations for the One-electron Atom" (PDF). University of Massachusetts Boston.
| date = 1926
P.J. Mohr, B.N. Taylor, and D.B. Newell (2011), "The 2010 CODATA Recommended Values of the Fundamental Physical Constants" (Web Version 6.0). This database was developed by J. Baker, M. Douma, and S. Kotochigova. Available: http://physics.nist.gov/constants. National Institute of Standards and Technology, Gaithersburg, MD 20899. Link to R∞, Link to hcR∞
| title = Über das Wasserstoffspektrum vom Standpunkt der neuen Quantenmechanik
Messiah, Albert (1999). Quantum Mechanics. New York: Dover. p. 1136. ISBN 0-486-40924-4.
| journal = Zeitschrift für Physik
LaguerreL. Wolfram Mathematica page
| volume = 36 | issue = 5 | pages = 336–363
Griffiths, p. 152
| doi = 10.1007/BF01450175
Condon and Shortley (1963). The Theory of Atomic Spectra. London: Cambridge. p. 441.
|bibcode = 1926ZPhy...36..336P }}</ref> using a rotational symmetry in four dimensions [O(4)-symmetry] generated by the angular momentum
Griffiths, Ch. 4 p. 89
and the [[Laplace–Runge–Lenz vector]]. By extending the symmetry group O(4) to the [[dynamical group]] O(4,2),
Bransden, B. H.; Joachain, C. J. (1983). Physics of Atoms and Molecules. Longman. p. Appendix 5. ISBN 0-582-44401-2.
the entire spectrum and all transitions were embedded in a single irreducible group representation.<ref>
Sommerfeld, Arnold (1919). Atombau und Spektrallinien'. Braunschweig: Friedrich Vieweg und Sohn. ISBN 3-87144-484-7. German English
{{cite journal
Summary of atomic quantum numbers. Lecture notes. 28 July 2006
| title = Group Dynamics of the Hydrogen Atom
Pauli, W (1926). "Über das Wasserstoffspektrum vom Standpunkt der neuen Quantenmechanik". Zeitschrift für Physik. 36 (5): 336–363. Bibcode:1926ZPhy...36..336P. doi:10.1007/BF01450175.
| author = Kleinert H.
Kleinert H. (1968). "Group Dynamics of the Hydrogen Atom" (PDF). Lectures in Theoretical Physics, edited by W.E. Brittin and A.O. Barut, Gordon and Breach, N.Y. 1968: 427–482.
| journal = Lectures in Theoretical Physics, edited by W.E. Brittin and A.O. Barut, Gordon and Breach, N.Y. 1968
Duru I.H., Kleinert H. (1979). "Solution of the path integral for the H-atom" (PDF). Physics Letters B. 84 (2): 185–188. Bibcode:1979PhLB...84..185D. doi:10.1016/0370-2693(79)90280-6.
| volume =
| pages = 427–482
| date = 1968
| doi =
| url = http://www.physik.fu-berlin.de/~kleinert/kleiner_re4/4.pdf
}}</ref>


In 1979 the (non relativistic) hydrogen atom was solved for the first time within [[R.P. Feynman|Feynman's]] [[path integral formulation]]
Duru I.H., Kleinert H. (1982). "Quantum Mechanics of H-Atom from Path Integrals" (PDF). Fortschr. Phys. 30 (2): 401–435. Bibcode:1982ForPh..30..401D. doi:10.1002/prop.19820300802.
of [[quantum mechanics]].<ref>
{{cite journal
| title = Solution of the path integral for the H-atom
| author = Duru I.H., Kleinert H.
| journal = Physics Letters B
| volume = 84
| issue = 2
| pages = 185–188
| date = 1979
| doi = 10.1016/0370-2693(79)90280-6
| url = http://www.physik.fu-berlin.de/~kleinert/kleiner_re65/65.pdf
|bibcode = 1979PhLB...84..185D }}</ref><ref>
{{cite journal
| title = Quantum Mechanics of H-Atom from Path Integrals
| author = Duru I.H., Kleinert H.
| journal = Fortschr. Phys
| volume = 30
| issue = 2
| pages = 401–435
| date = 1982
| doi = 10.1002/prop.19820300802
| url = http://www.physik.fu-berlin.de/~kleinert/kleiner_re83/83.pdf
|bibcode = 1982ForPh..30..401D }}</ref> This work greatly extended the range of applicability of [[R.P. Feynman|Feynman's]] method.


==See also==
Books
{{Wikipedia books|Hydrogen}}
{{col-begin|width=auto}}
{{col-break}}
* [[Antihydrogen]]
* [[Atomic orbital]]
* [[Balmer series]]
* [[Helium atom]]
* [[Lithium atom]]
* [[Dihydrogen cation|Hydrogen Molecular Ion]]
{{col-break}}
* [[Proton decay]]
* [[Quantum chemistry]]
* [[Quantum state]]
* [[Theoretical and experimental justification for the Schrödinger equation]]
* [[Trihydrogen cation]]
* [[List of quantum-mechanical systems with analytical solutions]]
{{col-end}}


==References==
Griffiths, David J. (1995). Introduction to Quantum Mechanics. Prentice Hall. ISBN 0-13-111892-7. Section 4.2 deals with the hydrogen atom specifically, but all of Chapter 4 is relevant.
{{reflist|30em}}
Kleinert, H. (2009). Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets, 4th edition, Worldscibooks.com, World Scientific, Singapore (also available online physik.fu-berlin.de)


==Books==
External links
*{{cite book
| first=David J.
| last=[[David J. Griffiths|Griffiths]]
| title=Introduction to Quantum Mechanics
| publisher=[[Prentice Hall]]
| date=1995
|isbn=0-13-111892-7
}} Section 4.2 deals with the hydrogen atom specifically, but all of Chapter 4 is relevant.
* [[Hagen Kleinert|Kleinert, H.]] (2009). ''Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets'', 4th edition, [http://www.worldscibooks.com/physics/7305.html Worldscibooks.com], World Scientific, Singapore (also available online [http://www.physik.fu-berlin.de/~kleinert/re.html#B8 physik.fu-berlin.de])


==External links==
Physics of hydrogen atom on Scienceworld
*[http://scienceworld.wolfram.com/physics/HydrogenAtom.html Physics of hydrogen atom on Scienceworld]
Applet which allows viewing of all sorts of hydrogenic orbitals
*[http://www.falstad.com/qmatom/ Applet which allows viewing of all sorts of hydrogenic orbitals]
Basic Quantum Mechanics of the Hydrogen Atom
*[http://www.physics.drexel.edu/~tim/open/hydrofin Basic Quantum Mechanics of the Hydrogen Atom]
"Research team takes image of hydrogen atom" Kyodo News, Friday, 5 November 2010 – (includes image)
*[http://search.japantimes.co.jp/cgi-bin/nn20101105a1.html "Research team takes image of hydrogen atom" Kyodo News, Friday, 5 November 2010 – (includes image)]


{{Isotope|element=hydrogen
|lighter=(none, lightest possible)
|heavier=[[Deuterium|hydrogen-2]]
|before=[[free neutron]]<br/>[[Diproton|helium-2]]
|after=Stable
}}


[[Category:Concepts in physics]]
Lighter:
[[Category:Atoms]]
(none, lightest possible) Hydrogen atom is an
[[Category:Quantum models]]
isotope of hydrogen Heavier:
[[Category:Hydrogen]]
hydrogen-2
[[Category:Hydrogen physics]]
Decay product of:
[[Category:Isotopes of hydrogen]]
free neutron
helium-2 Decay chain
of hydrogen atom Decays to:
Stable
Categories:


[[pl:Wodór atomowy]]
Concepts in physicsAtomsQuantum modelsHydrogenHydrogen physicsIsotopes of hydrogen

Revision as of 17:57, 6 May 2018

Hydrogen atom, 1H
General
Symbol1H
Nameshydrogen atom, 1H, H-1,
protium
Protons (Z)1
Neutrons (N)0
Nuclide data
Natural abundance99.985%
Isotope mass1.007825 Da
Spin1/2
Excess energy7288.969 keV
Binding energy0.000 keV
Isotopes of hydrogen
Complete table of nuclides
Depiction of a hydrogen atom showing the diameter as about twice the Bohr model radius. (Image not to scale)

A hydrogen atom is an atom of the chemical element hydrogen. The electrically neutral atom contains a single positively charged proton and a single negatively charged electron bound to the nucleus by the Coulomb force. Atomic hydrogen constitutes about 75% of the baryonic mass of the universe.[1]

In everyday life on Earth, isolated hydrogen atoms (called "atomic hydrogen") are extremely rare. Instead, hydrogen tends to combine with other atoms in compounds, or with itself to form ordinary (diatomic) hydrogen gas, H2. "Atomic hydrogen" and "hydrogen atom" in ordinary English use have overlapping, yet distinct, meanings. For example, a water molecule contains two hydrogen atoms, but does not contain atomic hydrogen (which would refer to isolated hydrogen atoms).

Attempts to develop a theoretical understanding of the hydrogen atom have been important to the history of quantum mechanics.This is the only element without a neutron

Isotopes

The most abundant isotope, hydrogen-1, protium, or light hydrogen, contains no neutrons and is simply a proton and an electron. Protium is stable and makes up 99.985% of naturally occurring hydrogen atoms.[2]

Deuterium contains one neutron and one proton. Deuterium is stable and makes up 0.0156% of naturally occurring hydrogen[2] and is used in industrial processes like nuclear reactors and Nuclear Magnetic Resonance.

Tritium contains two neutrons and one proton and is not stable, decaying with a half-life of 12.32 years. Because of the short half life, tritium does not exist in nature except in trace amounts.

Higher isotopes of hydrogen are only created in artificial accelerators and reactors and have half lives around the order of 10−22 seconds.

The formulas below are valid for all three isotopes of hydrogen, but slightly different values of the Rydberg constant (correction formula given below) must be used for each hydrogen isotope.

Hydrogen ion

Hydrogen is not found without its electron in ordinary chemistry (room temperatures and pressures), as ionized hydrogen is highly chemically reactive. When ionized hydrogen is written as "H+" as in the solvation of classical acids such as hydrochloric acid, the hydronium ion, H3O+, is meant, not a literal ionized single hydrogen atom. In that case, the acid transfers the proton to H2O to form H3O+.

Ionized hydrogen without its electron, or free protons, are common in the interstellar medium, and solar wind.

Theoretical analysis

The hydrogen atom has special significance in quantum mechanics and quantum field theory as a simple two-body problem physical system which has yielded many simple analytical solutions in closed-form.

Failed classical description

Experiments by Ernest Rutherford in 1909 showed the structure of the atom to be a dense, positive nucleus with a light, negative charge orbiting around it. This immediately caused problems on how such a system could be stable. Classical electromagnetism had shown that any accelerating charge radiates energy described through the Larmor formula. If the electron is assumed to orbit in a perfect circle and radiates energy continuously, the electron would rapidly spiral into the nucleus with a fall time of:[3]

Where is the Bohr radius and is the classical electron radius. If this were true, all atoms would instantly collapse, however atoms seem to be stable. Furthermore, the spiral inward would release a smear of electromagnetic frequencies as the orbit got smaller. Instead, atoms were observed to only emit discrete frequencies of radiation. The resolution would lie in the development of quantum mechanics.

Bohr-Sommerfeld Model

In 1913, Niels Bohr obtained the energy levels and spectral frequencies of the hydrogen atom after making a number of simple assumptions in order to correct the failed classical model. The assumptions included:

  1. Electrons can only be in certain, discrete circular orbits or stationary states, thereby having a discrete set of possible radii and energies.
  2. Electrons do not emit radiation while in one of these stationary states.
  3. An electron can gain or lose energy by jumping from one discrete orbital to another.

Bohr supposed that the electron's angular momentum is quantized with possible values:

where

and is Planck constant over . He also supposed that the centripetal force which keeps the electron in its orbit is provided by the Coulomb force, and that energy is conserved. Bohr derived the energy of each orbit of the hydrogen atom to be:[4]

,

where is the electron mass, is the electron charge, is the vacuum permittivity, and is the quantum number (now known as the principal quantum number). Bohr's predictions matched experiments measuring the hydrogen spectral series to the first order, giving more confidence to a theory that used quantized values.

For , the value

[5]

is called the Rydberg unit of energy. It is related to the Rydberg constant of atomic physics by

The exact value of the Rydberg constant assumes that the nucleus is infinitely massive with respect to the electron. For hydrogen-1, hydrogen-2 (deuterium), and hydrogen-3 (tritium) the constant must be slightly modified to use the reduced mass of the system, rather than simply the mass of the electron. However, since the nucleus is much heavier than the electron, the values are nearly the same. The Rydberg constant RM for a hydrogen atom (one electron), R is given by

where is the mass of the atomic nucleus. For hydrogen-1, the quantity is about 1/1836 (i.e. the electron-to-proton mass ratio). For deuterium and tritium, the ratios are about 1/3670 and 1/5497 respectively. These figures, when added to 1 in the denominator, represent very small corrections in the value of R, and thus only small corrections to all energy levels in corresponding hydrogen isotopes.

There were still problems with Bohr's model:

  1. it failed to predict other spectral details such as fine structure and hyperfine structure
  2. it could only predict energy levels with any accuracy for single–electron atoms (hydrogen–like atoms)
  3. the predicted values were only correct to , where is the fine-structure constant.

Most of these shortcomings were repaired by Arnold Sommerfeld's modification of the Bohr model. Sommerfeld introduced two additional degrees of freedom allowing an electron to move on an elliptical orbit, characterized by its eccentricity and declination with respect to a chosen axis. This introduces two additional quantum numbers, which correspond to the orbital angular momentum and its projection on the chosen axis. Thus the correct multiplicity of states (except for the factor 2 accounting for the yet unknown electron spin) was found. Further applying special relativity theory to the elliptic orbits, Sommerfeld succeeded in deriving the correct expression for the fine structure of hydrogen spectra (which happens to be exactly the same as in the most elaborate Dirac theory). However some observed phenomena such as the anomalous Zeeman effect remain unexplained. These issues were resolved with the full development of quantum mechanics and the Dirac equation. It is often alleged, that the Schrödinger equation is superior to the Bohr-Sommerfeld theory in describing hydrogen atom. This is however not the case, as the most results of both approaches coincide or are very close (a remarkable exception is the problem of hydrogen atom in crossed electric and magnetic fields, which cannot be solved in the framework of the Bohr-Sommerfeld theory self-consistently), and their main shortcomings result from the absence of the electron spin in both theories. It was the complete failure of the Bohr-Sommerfeld theory to explain many-electron systems (such as helium atom or hydrogen molecule) which demonstrated its inadequacy in describing quantum phenomena.

Schrödinger equation

The Schrödinger equation allows one to calculate the development of quantum systems with time and can give exact, analytical answers for the non-relativistic hydrogen atom.

Wavefunction

The Hamiltonian of the hydrogen atom is the radial kinetic energy operator and coulomb attraction force between the positive proton and negative electron. Using the time-independent Schrödinger equation, ignoring all spin-coupling interactions and using the reduced mass , the equation is written as:

Expanding the Laplacian in spherical coordinates:

This is a separable, partial differential equation which can be solved in terms of special functions. The normalized position wavefunctions, given in spherical coordinates are:

3D illustration of the eigenstate . Electrons in this state are 45% likely to be found within the solid body shown.

where:

,
is the reduced Bohr radius, ,
is a generalized Laguerre polynomial of degree n − 1, and
is a spherical harmonic function of degree and order m. Note that the generalized Laguerre polynomials are defined differently by different authors. The usage here is consistent with the definitions used by Messiah,[6] and Mathematica.[7] In other places, the Laguerre polynomial includes a factor of ,[8] or the generalized Laguerre polynomial appearing in the hydrogen wave function is instead.[9]

The quantum numbers can take the following values:

Additionally, these wavefunctions are normalized (i.e., the integral of their modulus square equals 1) and orthogonal:

where is the state represented by the wavefunction in Dirac notation, and is the Kronecker delta function.[10]

The wavefunctions in momentum space are related to the wavefunctions in position space through a Fourier transform

which, for the bound states, results in [11]

where denotes a Gegenbauer polynomial and is in units of .

The solutions to the Schrödinger equation for hydrogen are analytical, giving a simple expression for the hydrogen energy levels and thus the frequencies of the hydrogen spectral lines and fully reproduced the Bohr model and went beyond it. It also yields two other quantum numbers and the shape of the electron's wave function ("orbital") for the various possible quantum-mechanical states, thus explaining the anisotropic character of atomic bonds.

The Schrödinger equation also applies to more complicated atoms and molecules. When there is more than one electron or nucleus the solution is not analytical and either computer calculations are necessary or simplifying assumptions must be made.

Since the Schrödinger equation is only valid for non-relativistic quantum mechanics, the solutions it yields for the hydrogen atom are not entirely correct. The Dirac equation of relativistic quantum theory improves these solutions (see below).

Results of Schrödinger equation

The solution of the Schrödinger equation (wave equation) for the hydrogen atom uses the fact that the Coulomb potential produced by the nucleus is isotropic (it is radially symmetric in space and only depends on the distance to the nucleus). Although the resulting energy eigenfunctions (the orbitals) are not necessarily isotropic themselves, their dependence on the angular coordinates follows completely generally from this isotropy of the underlying potential: the eigenstates of the Hamiltonian (that is, the energy eigenstates) can be chosen as simultaneous eigenstates of the angular momentum operator. This corresponds to the fact that angular momentum is conserved in the orbital motion of the electron around the nucleus. Therefore, the energy eigenstates may be classified by two angular momentum quantum numbers, and m (both are integers). The angular momentum quantum number = 0, 1, 2, ... determines the magnitude of the angular momentum. The magnetic quantum number m = −, ..., + determines the projection of the angular momentum on the (arbitrarily chosen) z-axis.

In addition to mathematical expressions for total angular momentum and angular momentum projection of wavefunctions, an expression for the radial dependence of the wave functions must be found. It is only here that the details of the 1/r Coulomb potential enter (leading to Laguerre polynomials in r). This leads to a third quantum number, the principal quantum number n = 1, 2, 3, .... The principal quantum number in hydrogen is related to the atom's total energy.

Note that the maximum value of the angular momentum quantum number is limited by the principal quantum number: it can run only up to n − 1, i.e. = 0, 1, ..., n − 1.

Due to angular momentum conservation, states of the same but different m have the same energy (this holds for all problems with rotational symmetry). In addition, for the hydrogen atom, states of the same n but different are also degenerate (i.e. they have the same energy). However, this is a specific property of hydrogen and is no longer true for more complicated atoms which have an (effective) potential differing from the form 1/r (due to the presence of the inner electrons shielding the nucleus potential).

Taking into account the spin of the electron adds a last quantum number, the projection of the electron's spin angular momentum along the z-axis, which can take on two values. Therefore, any eigenstate of the electron in the hydrogen atom is described fully by four quantum numbers. According to the usual rules of quantum mechanics, the actual state of the electron may be any superposition of these states. This explains also why the choice of z-axis for the directional quantization of the angular momentum vector is immaterial: an orbital of given and m′ obtained for another preferred axis z′ can always be represented as a suitable superposition of the various states of different m (but same l) that have been obtained for z.

Mathematical summary of eigenstates of hydrogen atom

In 1928, Paul Dirac found an equation that was fully compatible with Special Relativity, and (as a consequence) made the wave function a 4-component "Dirac spinor" including "up" and "down" spin components, with both positive and "negative" energy (or matter and antimatter). The solution to this equation gave the following results, more accurate than the Schrödinger solution.

Energy levels

The energy levels of hydrogen, including fine structure (excluding Lamb shift and hyperfine structure), are given by the Sommerfeld fine structure expression:[12]

where α is the fine-structure constant and j is the "total angular momentum" quantum number, which is equal to | ± 1/2| depending on the direction of the electron spin. This formula represents a small correction to the energy obtained by Bohr and Schrödinger as given above. The factor in square brackets in the last expression is nearly one; the extra term arises from relativistic effects (for details, see #Features going beyond the Schrödinger solution). It is worth noting that this expression was first obtained by A. Sommerfeld in 1916 based on the relativistic version of the old Bohr theory. Sommerfeld has however used different notation for the quantum numbers.

Visualizing the hydrogen electron orbitals

Probability densities through the xz-plane for the electron at different quantum numbers (, across top; n, down side; m = 0)

The image to the right shows the first few hydrogen atom orbitals (energy eigenfunctions). These are cross-sections of the probability density that are color-coded (black represents zero density and white represents the highest density). The angular momentum (orbital) quantum number is denoted in each column, using the usual spectroscopic letter code (s means  = 0, p means  = 1, d means  = 2). The main (principal) quantum number n (= 1, 2, 3, ...) is marked to the right of each row. For all pictures the magnetic quantum number m has been set to 0, and the cross-sectional plane is the xz-plane (z is the vertical axis). The probability density in three-dimensional space is obtained by rotating the one shown here around the z-axis.

The "ground state", i.e. the state of lowest energy, in which the electron is usually found, is the first one, the 1s state (principal quantum level n = 1, = 0).

Black lines occur in each but the first orbital: these are the nodes of the wavefunction, i.e. where the probability density is zero. (More precisely, the nodes are spherical harmonics that appear as a result of solving Schrödinger equation in polar coordinates.)

The quantum numbers determine the layout of these nodes.[13] There are:

  • total nodes,
  • of which are angular nodes:
    • angular nodes go around the axis (in the xy plane). (The figure above does not show these nodes since it plots cross-sections through the xz-plane.)
    • (the remaining angular nodes) occur on the (vertical) axis.
  • (the remaining non-angular nodes) are radial nodes.

Features going beyond the Schrödinger solution

There are several important effects that are neglected by the Schrödinger equation and which are responsible for certain small but measurable deviations of the real spectral lines from the predicted ones:

  • Although the mean speed of the electron in hydrogen is only 1/137th of the speed of light, many modern experiments are sufficiently precise that a complete theoretical explanation requires a fully relativistic treatment of the problem. A relativistic treatment results in a momentum increase of about 1 part in 37,000 for the electron. Since the electron's wavelength is determined by its momentum, orbitals containing higher speed electrons show contraction due to smaller wavelengths.
  • Even when there is no external magnetic field, in the inertial frame of the moving electron, the electromagnetic field of the nucleus has a magnetic component. The spin of the electron has an associated magnetic moment which interacts with this magnetic field. This effect is also explained by special relativity, and it leads to the so-called spin-orbit coupling, i.e., an interaction between the electron's orbital motion around the nucleus, and its spin.

Both of these features (and more) are incorporated in the relativistic Dirac equation, with predictions that come still closer to experiment. Again the Dirac equation may be solved analytically in the special case of a two-body system, such as the hydrogen atom. The resulting solution quantum states now must be classified by the total angular momentum number j (arising through the coupling between electron spin and orbital angular momentum). States of the same j and the same n are still degenerate. Thus, direct analytical solution of Dirac equation predicts 2S(1/2) and 2P(1/2) levels of Hydrogen to have exactly the same energy, which is in a contradiction with observations (Lamb-Retherford experiment).

For these developments, it was essential that the solution of the Dirac equation for the hydrogen atom could be worked out exactly, such that any experimentally observed deviation had to be taken seriously as a signal of failure of the theory.

Alternatives to the Schrödinger theory

In the language of Heisenberg's matrix mechanics, the hydrogen atom was first solved by Wolfgang Pauli[14] using a rotational symmetry in four dimensions [O(4)-symmetry] generated by the angular momentum and the Laplace–Runge–Lenz vector. By extending the symmetry group O(4) to the dynamical group O(4,2), the entire spectrum and all transitions were embedded in a single irreducible group representation.[15]

In 1979 the (non relativistic) hydrogen atom was solved for the first time within Feynman's path integral formulation of quantum mechanics.[16][17] This work greatly extended the range of applicability of Feynman's method.

See also

Template:Wikipedia books

References

  1. ^ Palmer, D. (13 September 1997). "Hydrogen in the Universe". NASA. Archived from the original on 29 October 2014. Retrieved 23 February 2017. {{cite web}}: Unknown parameter |dead-url= ignored (|url-status= suggested) (help)
  2. ^ a b Housecroft, Catherine E.; Sharpe, Alan G. (2005). Inorganic Chemistry (2nd ed.). Pearson Prentice-Hall. p. 237. ISBN 0130-39913-2.
  3. ^ Olsen, James; McDonald, Kirk (7 March 2005). "Classical Lifetime of a Bohr Atom" (PDF). Joseph Henry Laboratories, Princeton University.
  4. ^ "Derivation of Bohr's Equations for the One-electron Atom" (PDF). University of Massachusetts Boston.
  5. ^ P.J. Mohr, B.N. Taylor, and D.B. Newell (2011), "The 2010 CODATA Recommended Values of the Fundamental Physical Constants" (Web Version 6.0). This database was developed by J. Baker, M. Douma, and S. Kotochigova. Available: http://physics.nist.gov/constants. National Institute of Standards and Technology, Gaithersburg, MD 20899. Link to R, Link to hcR
  6. ^ Messiah, Albert (1999). Quantum Mechanics. New York: Dover. p. 1136. ISBN 0-486-40924-4.
  7. ^ LaguerreL. Wolfram Mathematica page
  8. ^ Griffiths, p. 152
  9. ^ Condon and Shortley (1963). The Theory of Atomic Spectra. London: Cambridge. p. 441.
  10. ^ Griffiths, Ch. 4 p. 89
  11. ^ Bransden, B. H.; Joachain, C. J. (1983). Physics of Atoms and Molecules. Longman. p. Appendix 5. ISBN 0-582-44401-2.
  12. ^ Sommerfeld, Arnold (1919). Atombau und Spektrallinien'. Braunschweig: Friedrich Vieweg und Sohn. ISBN 3-87144-484-7. German English
  13. ^ Summary of atomic quantum numbers. Lecture notes. 28 July 2006
  14. ^ Pauli, W (1926). "Über das Wasserstoffspektrum vom Standpunkt der neuen Quantenmechanik". Zeitschrift für Physik. 36 (5): 336–363. Bibcode:1926ZPhy...36..336P. doi:10.1007/BF01450175.
  15. ^ Kleinert H. (1968). "Group Dynamics of the Hydrogen Atom" (PDF). Lectures in Theoretical Physics, edited by W.E. Brittin and A.O. Barut, Gordon and Breach, N.Y. 1968: 427–482.
  16. ^ Duru I.H., Kleinert H. (1979). "Solution of the path integral for the H-atom" (PDF). Physics Letters B. 84 (2): 185–188. Bibcode:1979PhLB...84..185D. doi:10.1016/0370-2693(79)90280-6.
  17. ^ Duru I.H., Kleinert H. (1982). "Quantum Mechanics of H-Atom from Path Integrals" (PDF). Fortschr. Phys. 30 (2): 401–435. Bibcode:1982ForPh..30..401D. doi:10.1002/prop.19820300802.

Books


Lighter:
(none, lightest possible)
Hydrogen atom is an
isotope of hydrogen
Heavier:
hydrogen-2
Decay product of:
free neutron
helium-2
Decay chain
of hydrogen atom
Decays to:
Stable